Majorana Fermion Surface Code for Universal Quantum Computation Please share

advertisement
Majorana Fermion Surface Code for Universal Quantum
Computation
The MIT Faculty has made this article openly available. Please share
how this access benefits you. Your story matters.
Citation
Vijay, Sagar, Timothy H. Hsieh, and Liang Fu. "Majorana
Fermion Surface Code for Universal Quantum Computation."
Phys. Rev. X 5, 041038 (December 2015).
As Published
http://dx.doi.org/10.1103/PhysRevX.5.041038
Publisher
American Physical Society
Version
Final published version
Accessed
Thu May 26 18:50:21 EDT 2016
Citable Link
http://hdl.handle.net/1721.1/100455
Terms of Use
Creative Commons Attribution
Detailed Terms
http://creativecommons.org/licenses/by/3.0
PHYSICAL REVIEW X 5, 041038 (2015)
Majorana Fermion Surface Code for Universal Quantum Computation
Sagar Vijay, Timothy H. Hsieh, and Liang Fu
Department of Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA
(Received 8 April 2015; published 10 December 2015)
We introduce an exactly solvable model of interacting Majorana fermions realizing Z2 topological order
with a Z2 fermion parity grading and lattice symmetries permuting the three fundamental anyon types. We
propose a concrete physical realization by utilizing quantum phase slips in an array of Josephson-coupled
mesoscopic topological superconductors, which can be implemented in a wide range of solid-state systems,
including topological insulators, nanowires, or two-dimensional electron gases, proximitized by s-wave
superconductors. Our model finds a natural application as a Majorana fermion surface code for universal
quantum computation, with a single-step stabilizer measurement requiring no physical ancilla qubits,
increased error tolerance, and simpler logical gates than a surface code with bosonic physical qubits.
We thoroughly discuss protocols for stabilizer measurements, encoding and manipulating logical
qubits, and gate implementations.
DOI: 10.1103/PhysRevX.5.041038
Subject Areas: Condensed Matter Physics,
Quantum Information,
Superconductivity
I. INTRODUCTION
As originally proposed by Ettore Majorana, the
Majorana fermion is a particle that is its own antiparticle
[1]. In the condensed-matter setting, Majorana fermions
can emerge in topological superconductors as a special type
of zero-energy quasiparticle that is bound to vortices or
defects and formed by an equal superposition of electron
and hole excitations [2,3]. Theory predicts that Majorana
fermions can be created in a wide array of spinorbit-coupled materials in proximity to conventional superconductors [4–8]. Recent observations of zero-energy
conductance peaks in such systems [9–12] provide encouraging hints of Majorana fermions [13,14].
Majorana fermions in topological superconductors are of
great interest as they are predicted to produce exotic
quantum phenomena such as the fractional Josephson effect
[15–17] and electron teleportation [18]. Most remarkably,
vortices or defects that carry Majorana fermions are predicted to exhibit non-Abelian statistics [19–21], which have
yet to be observed in nature. In addition to its theoretical
significance, non-Abelian statistics provides the foundation
for topological quantum computation, in which logical
qubits are encoded in the topologically degenerate states
of non-Abelian anyons, and qubit operations are performed
by braiding [22]. Topological quantum computation has the
theoretical advantage of being immune to errors caused by
local perturbations [23]. Demonstrating the non-Abelian
Published by the American Physical Society under the terms of
the Creative Commons Attribution 3.0 License. Further distribution of this work must maintain attribution to the author(s) and
the published article’s title, journal citation, and DOI.
2160-3308=15=5(4)=041038(17)
statistics of Majorana fermions, however, requires braiding,
fusing, and measuring the fusion outcome. This is a
challenging task, as each of the above operations is yet to
be experimentally achieved. Furthermore, braiding Majorana
fermions alone is insufficient to perform the necessary gate
operations for universal quantum computation.
The “surface code” [24,25] provides an alternative
approach to universal quantum computation that uses
measurements in an Abelian topological phase for gate
operations and error correction. In the surface code,
measurements of nontrivial commuting operators (stabilizers) are used to project onto a “code state,” and logical
qubits are effectively encoded in the anyon charge of a
region by ceasing certain stabilizer measurements [26–29].
The logical gates necessary for universal quantum computation are realized through sequences of measurements used
to move and braid the logical qubits. An advantage of the
surface code architecture is its remarkable ability for error
detection and subsequent correction during qubit readout,
as the nucleation of anyons through the action of a random
operator can be reliably tracked through stabilizer measurements. For a sufficiently low error rate per physical
qubit measurement, scaling the size of the surface code
produces an exponential suppression in propagated errors
[30]. Remarkably, recent experiments with superconducting quantum circuits have demonstrated the ability to
perform high-fidelity physical gate operations and reliable
error correction for a surface code of small size [31–33].
In this work, we introduce a new scheme for surface code
quantum computation that uses Majorana fermions as the
fundamental physical degrees of freedom and exploits their
unique properties for encoding and manipulating logical
qubits. Our surface code is based on a novel Z2 topological
041038-1
Published by the American Physical Society
SAGAR VIJAY, TIMOTHY H. HSIEH, AND LIANG FU
order with fermion parity grading (defined below), which
we demonstrate in a class of exactly solvable Hamiltonians
of interacting Majorana fermions, termed Majorana plaquette models. We demonstrate that charging energyinduced quantum phase slips in superconducting arrays
with Majorana fermions generates the required multifermion plaquette interactions, providing a physical realization of our model. We then describe a detailed physical
implementation of the “Majorana fermion surface code,”
including physical qubit and stabilizer measurements, the
creation of logical qubits, error correction, and logical gate
operations required for universal quantum computation.
The Majorana fermion surface code poses significant
benefits over a surface code with bosonic physical qubits.
First, stabilizer measurements in the Majorana surface code
can be performed in a single step, whereas this requires
several physical gate operations in the conventional surface
code [28,29]. As a result, we anticipate that the Majorana
surface code has a significantly higher error tolerance.
Furthermore, our Majorana surface code operates with
substantially less overhead, as it requires fewer physical
qubits per encoded logical qubit and uses no physical
ancilla qubits. Second, we may tune the energy gap for
anyon excitations in our physical realization of the
Majorana plaquette Hamiltonian, increasing error suppression in the Majorana fermion surface code. Finally, the
lattice symmetries in the Majorana plaquette model permute the three fundamental anyon types, allowing for a
much simpler implementation of the logical Hadamard
gate. As we will show, these advantages arise from the
unique approach taken by our Majorana fermion surface
code and the use of Majorana fermions as the fundamental
degrees of freedom. The fact that a Majorana fermion is
“half” of an ordinary fermion—so that a pair is required to
encode a qubit of information—is crucial to the Majorana
fermion surface code. We emphasize that the non-Abelian
statistics of Majorana-carrying vortices or defects is of no
relevance to our proposal, as our code does not involve
braiding them.
This paper is organized as follows. First, we introduce a
solvable model of interacting Majorana fermions on the
honeycomb lattice realizing a novel Z2 topological order
with a Z2 fermion parity grading and an exact S3 anyon
symmetry. We propose a physical realization of this model
that uses the charging energy in an array of mesoscopic
superconductors [18] to implement the required nonlocal
interactions between multiple Majorana fermions. Next, we
demonstrate that our model provides a natural setting for
the Majorana fermion surface code, in which a logical qubit
is encoded in a set of physical qubits formed from
Majorana fermions. We present a physical implementation
of the Majorana surface code and propose detailed protocols for performing gate operations for universal quantum
computation.
PHYS. REV. X 5, 041038 (2015)
FIG. 1. We consider a honeycomb lattice with (a) a single
Majorana fermion on each lattice site, so the Op operator is the
product of the six Majorana fermions on the vertices of a
hexagonal plaquette. The colored plaquettes in (b) correspond
to the three distinct bosonic excitations that may be obtained by
violating a plaquette constraint.
II. MAJORANA PLAQUETTE MODEL
We begin by considering a honeycomb lattice with one
Majorana fermion (γ) on each lattice site; the Majorana
fermions satisfy canonical anticommutation relations
fγ n ; γ m g ¼ 2δnm . The Hamiltonian is defined as the sum
of operators acting on each hexagonal plaquette:
X
H ¼ −u Op ;
Op ≡ i
p
Y
γn:
ð1Þ
n∈vertexðpÞ
We note that this model was mentioned in a work by
Bravyi, Terhal, and Leemhuis [34], although its novel
topological order and anyon excitations were not studied
there; a closely related model in the same topological phase
was introduced and studied by Xu and Fu [35].
It suffices to consider u > 0 in the Hamiltonian (1), as
the case with u < 0 can be mapped to u > 0 by changing
the sign of the Majorana fermions on one sublattice.
The operator Op is the product of the six Majorana
fermions on the vertices of plaquette p as shown in
Fig. 1(a). Since any two plaquettes on the honeycomb
lattice share an even number of vertices, all of the
plaquette operators commute, and the ground-state jΨ0 i
is defined by the condition
Op jΨ0 i ¼ jΨ0 i;
ð2Þ
for all plaquettes p. We note that, quite generally,
Hamiltonians of interacting Majorana fermions with
commuting terms may be realized on any lattice, so long
as any pair of operators in the Hamiltonian only has
overlapping support over an even number of Majorana
fermions.
We demonstrate that the above Majorana plaquette
model (1) realizes a Z2 topological order of Fermi systems
by considering the ground-state degeneracy and elementary
excitations. First, we place the system on a torus by
041038-2
MAJORANA FERMION SURFACE CODE FOR UNIVERSAL …
PHYS. REV. X 5, 041038 (2015)
imposing periodic boundary conditions and find a four-fold
degenerate ground state by counting the number of degrees
of freedom and constraints on the full Hilbert space. For an
N-site honeycomb lattice, the 2N=2 -dimensional Hilbert
space of Majorana fermions is constrained by the fixed total
fermion parity:
Γ ≡ iN=2
Y
γn:
ð3Þ
n
For convenience, we choose a unit cell for the honeycomb
lattice consisting of three plaquettes labeled A, B, and C,
as shown in Fig. 1(b). We observe that on the torus, the
product of plaquette operators on each of the A-, B-, and
C-type plaquettes is equal to the total fermion parity:
Γ¼
Y
Op ¼
p∈A
Y
Op ¼
p∈B
Y
Op :
ð4Þ
p∈C
The operators fOp g on any one type of plaquette fix
one-third of the plaquette eigenvalues via the condition (2)
and impose 2N=6−1 constraints on the Hilbert space. The
number of unconstrained degrees of freedom is therefore
given by
D ¼ 2ðN=2Þ−1 =ð2ðN=6Þ−1 Þ3 ¼ 4;
ð5Þ
which yields a four-fold ground-state degeneracy for the
Majorana plaquette model on the torus.
The ground-state degeneracy is of a topological nature,
as the four ground states are distinguished only by nonlocal
operators. To see this, we construct a Wilson loop operator
W l , defined as a product of Majorana bilinears on a
noncontractible loop l on the torus:
Wl ≡
Y
ðiγ n γ m Þ;
ð6Þ
n;m∈l
such that W 2l ¼ 1 so that the Wilson loop has eigenvalues
1. Consider the operators W x and W y on the two
nontrivial cycles of the torus lx and ly , as shown in
Fig. 2. Since lx and ly traverse an even number of
vertices over any plaquette and do not contain any common
lattice sites, we have ½W x ; W y ¼ ½W x ; H ¼ ½W y ; H ¼ 0.
Furthermore, we may construct Wilson loop operators W x~
and W y~ on loops l~ x and l~ y , where l~ x is shifted from lx by a
basis vector parallel to ly and likewise for l~ y, such that
fW x~ ; W y g ¼ fW y~ ; W x g ¼ 0. As before, W x~ and W y~ commute with each other and with the Hamiltonian. Therefore,
the four degenerate ground states may be distinguished by
their eigenvalues under W x and W y , with W x~ and W y~
transforming the ground states between distinct sectors. In
analogy with conventional Z2 gauge theory, we may
identify the Wilson loop operators W x;y with electric
FIG. 2. The action of the commuting Wilson loop operators W x
and W y is shown above as the product of the Majorana fermions
on the lattice sites intersected by the appropriate colored lines.
The operator W y~ anticommutes with W x and takes the ground
state between two topological sectors.
charges traversing the torus in two different directions,
and W x~ ;~y as magnetic fluxes on a dual lattice.
Gapped excitations above the ground state are obtained
by flipping the eigenvalue of Op from þ1 to −1 on one or
more plaquettes. Since the total fermion parity is fixed
and equal to the product of all plaquette operators of each
type, plaquette eigenvalues can only be flipped on pairs of
plaquettes of the same type. This is achieved by string
operators of the form (6), now acting on open paths
and anticommuting with the plaquette operators at the
two ends of the path, thereby creating a pair of anyon
excitations.
An important feature of our Majorana plaquette model,
the conservation of total fermion parity—a universal
property of Fermi systems—makes it impossible to create
or annihilate two excitations living on different types of
plaquettes or to change one type of plaquette excitation into
another. As a result, there are three distinct elementary
plaquette excitations, labeled A, B, and C, by plaquette
type. To determine their statistics, we braid these excitations by acting with Majorana hopping operators iγ n γ m on
lattice bonds [36]. We find that all three types of plaquette
excitations have boson self-statistics and mutual semion
statistics, i.e., braiding two distinct plaquette excitations
generates a quantized Berry phase of π. From the elementary plaquette excitations, we may build composite excitations AB, BC, AC, and ABC by flipping the eigenvalues
of the Op ’s on two or three adjacent plaquettes. Among
these, the composite excitation ABC is simply a physical
Majorana fermion since the Majorana operator γ n acting on
a lattice site flips the eigenvalues of the Op ’s on the three
surrounding A, B, and C plaquettes. In contrast, the
composite excitations AB, BC, and AC are anyons, with
fermion self-statistics and mutual semion statistics with the
elementary excitations. We call these excitations composite
Majorana fermions, as they are created by a string of
physical Majorana fermions. A summary of the braiding
statistics for all anyons in our Majorana plaquette model is
given in the following table:
041038-3
SAGAR VIJAY, TIMOTHY H. HSIEH, AND LIANG FU
1
A
B
C
AB
BC
AC
ABC
1
A
B
C
AB
BC
AC
ABC
þ1
þ1
þ1
þ1
þ1
þ1
þ1
þ1
þ1
þ1
−1
−1
−1
þ1
−1
þ1
þ1
−1
þ1
−1
−1
−1
þ1
þ1
þ1
−1
−1
þ1
þ1
−1
−1
þ1
þ1
−1
−1
þ1
−1
−1
−1
þ1
þ1
þ1
−1
−1
−1
−1
−1
þ1
þ1
−1
þ1
−1
−1
−1
−1
þ1
þ1
þ1
þ1
þ1
þ1
þ1
þ1
−1
Strange as it may seem, the existence of eight types of
quasiparticle excitations is a generic property of Z2 topologically ordered phases in Fermi systems, due to the
conservation of fermion parity. Consider artificially dividing
the above quasiparticles into two groups: ð1; A; B; ABÞ and
ðABC; BC; AC; CÞ ¼ ABC × ð1; A; B; ABÞ. The former is
equivalent to the four quasiparticles in Z2 gauge theory
coupled to a bosonic Ising matter field, as realized in
Kitaev’s toric code [22] or Wen’s plaquette model [37].
The latter group of quasiparticles is obtained by attaching a
physical Majorana fermion to the former. The conservation
of total fermion parity guarantees that the two groups of
quasiparticles cannot transform into each other in a closed
system and thus have separate identities. We refer to the
presence of two groups of excitations with different fermion
parity as a Z2 fermion parity grading.
A remarkable property of the Majorana plaquette model
is that crystal symmetries of the honeycomb lattice permute
the three fundamental anyon excitations, A, B, and C, by
interchanging the three types of plaquettes. Examples of
such lattice symmetries include π=3 rotations about the
center of a plaquette and translation by any primitive lattice
vector. These symmetries of the honeycomb lattice provide
a microscopic realization of the S3 anyon symmetry that
permutes quasiparticle sectors, as recently studied in the
formalism of topological field theory by considering the
symmetries of the K matrices of Abelian topological
states [38,39].
III. PHYSICAL REALIZATION
A. Physical platforms
In this section, we show that the Majorana plaquette
model can be physically realized in an array of mesoscopic
topological superconductors that are Josephson coupled. A
wide range of material platforms for engineering a topological superconductor have been proposed and are being
experimentally studied [2,3]. As will be clear in the
following, the scheme we propose for realizing the
Majorana plaquette model is independent of which platform is used. For the sake of concreteness, we use a
platform based on topological insulators in describing the
general scheme below and discuss other platforms based on
nanowires and a two-dimensional electron gas with spinorbit coupling in Sec. III C.
PHYS. REV. X 5, 041038 (2015)
FIG. 3. Array of hexagonal s-wave superconducting islands
placed on a TI surface. Each arrow points in the direction of the
relative phase of the associated island, with φ ¼ 0, 2π=3. This
produces a honeycomb lattice of vortices (blue) and antivortices
(red) at trijunctions, hosting Majorana fermions.
We place an array of hexagon-shaped s-wave superconducting islands on a topological insulator (TI) to induce
a superconducting proximity effect on the TI surface states.
The Hamiltonian for this superconductor-TI hybrid system
is given by
Z
H0 ¼
drð−ivÞψ † ðrÞð∂ x sy − ∂ y sx − μÞψðrÞ
XZ
þ
drj ½Δeiφj ψ †↑ ðrj Þψ †↓ ðrj Þ þ H:c:;
ð7Þ
j
where ψ ¼ ðψ ↑ ; ψ ↓ ÞT is a two-component fermion field
and sx;y are spin Pauli matrices. The first term describes the
pristine TI surface states, with a single spin-nondegenerate
Fermi surface and helical spin texture in momentum space.
The second term describes the superconducting proximity
effect: rj belongs to the region underneath the jth superconducting island, whose phase is denoted by φj.
As found by Fu and Kane [4], a vortex or antivortex
trapped at a trijunction, where three islands meet, hosts a
single Majorana fermion zero mode. Let us consider setting
up the phases of superconducting islands to realize an array
of vortices and antivortices at trijunctions. For example, the
phases can be set to φj ¼ 0, 2π=3, and −2π=3 on the A-,
B-, and C-type islands, respectively, as shown in Fig. 3.
This yields a two-dimensional (2D) array of Majorana
fermions on a honeycomb lattice. In practice, the desired
phase configuration can be engineered by external electrical circuits [40] and/or magnetic flux. Alternatively,
applying a perpendicular magnetic field generates a vortex
lattice. These vortices may naturally sit at these trijunctions
where the induced superconductivity is weak, leading to the
desired lattice of Majorana fermions.
We take the size of the islands to be larger than the
coherence length of the superconducting TI surface states.
Under this condition, Majorana fermions at different sites
have negligible wave-function overlap, preventing any
041038-4
MAJORANA FERMION SURFACE CODE FOR UNIVERSAL …
unwanted direct coupling between them. (We note that even
weak couplings from wave-function overlap will not affect
the Z2 topological order of the Majorana plaquette model
because of its finite energy gap.) Nonetheless, as we show
below, the charging energy of superconductors induces a
nonlocal interaction between the six Majorana fermions on
each island, providing the key ingredient of the Majorana
plaquette model.
B. Phase-slip-induced multifermion interactions
The important but subtle interplay between Majorana
fermions and charging energy was first recognized by Fu
and formulated for superconductors with a fixed number of
electrons [18]. Later works have extended this to multiple
superconductors connected by Josephson coupling and
single-electron tunneling [35,41–43]. In all of these cases,
the charging energy of a given superconductor induces
quantum phase slips φ → φ 2π, from which the
Majorana fermions in the superconductor acquire a minus
sign: γ i → −γ i . This property is due to the inherently
double-valued dependence of Majorana operators on the
superconducting phase [18].
In our setup for the Majorana plaquette model, the
charging energy of the superconducting islands exerts
even more dramatic and interesting effects on the
Majorana fermions at trijunctions, which have not been
previously studied. In the presence of a charging energy, the
phase of each island becomes a quantum rotor. The kinetic
energy of the rotor is provided by the charging energy Ec,
which depends on the capacitance between an island and
the rest of the array and is described by the following
Hamiltonian:
Hc ¼ 4Ec
X
ðn̂j − ng Þ2 ;
ð8Þ
We work in the regime EJ ≫ Ec . Under this condition, lowenergy states of the quantum rotor on a given island φj
consist of small-amplitude fluctuations around each potential minimum φ0;j þ 2πm. Moreover, different minima are
connected by quantum phase slips, in which the phase φ
tunnels through a high-energy barrier to wind by 2πn, with
n an integer. The small-amplitude phase fluctuations
around a potential minimum correspond to a quantum
harmonic oscillator and thus generate a set of energy levels
given by
pffiffiffiffiffiffiffiffiffiffiffiffiffi
ϵ0α ≈ ðα þ 1=2Þ 8EJ Ec ;
HJ ¼ −EJ
X
cosðφ̂j − φ̂j0 − ajj0 Þ;
ð9Þ
hj;j0 i
where ajj0 ¼ φ0;j − φ0;j0 is externally set up such that the
minimum of the Josephson energy corresponds to φj ¼
φ0;j mod 2π, with φ0;j ¼ 0, 2π=3, and −2π=3 for the A, B,
and C-type islands, respectively.
Combining Eqs. (7), (8), and (9), the full Hamiltonian for
our setup, i.e., an array of superconducting islands on a TI
surface, is given by
H ¼ H0 þ Hc þ HJ :
ð10Þ
ð11Þ
with α ∈ N.
On the other hand, quantum phase slips on a superconducting island strongly couple to the Majorana fermions
that reside on the border with its neighbors, previously
obtained by holding the phase fixed at φ0;j . In other words,
Majorana fermions enter the low-energy effective theory of
Eq. (10) via quantum phase slips induced by the small
charging energy on each superconducting island. This new
physics makes our system different from a conventional
Cooper-pair box. Remarkably, the action of a quantum
phase slip involves Majorana fermions in a way that
depends periodically on the phase winding number n
mod 6. Consider, for example, phase slips at the central
superconducting island in Fig. 3. For n ¼ 1, a 2π phase slip
φ ¼ 0 → 2π cyclically permutes the three Majorana fermions bound to vortices in the counterclockwise direction
and the three Majorana fermions bound to antivortices in
the clockwise direction, i.e.,
φ ¼ 0 → 2π∶ γ 1 → γ 3 ;
γ 2 → −γ 6 ;
j
where n̂j ≡ ð−iÞ∂=∂φj is the Cooper-pair number operator
for the jth island and ng is the offset charge, which can be
tuned by an externally applied electric field. The potential
energy of the rotor is provided by the Josephson coupling
EJ between adjacent superconducting islands, given by
PHYS. REV. X 5, 041038 (2015)
γ3 → γ5 ;
γ4 → γ2 ;
γ 5 → −γ 1
γ6 → γ4;
ð12Þ
as shown in Fig. 4, where i ¼ 1; …; 6 labels the six
Majorana fermions at vertices of this island in clockwise
order. The physical movement of Majorana fermions
induced by phase slips is a unique and attractive advantage
of our setup, compared to other setups in which the
positions of Majorana fermions are fixed [35,41–43]. On
the other hand, for n ¼ 3, a 6π phase slip takes each
Majorana fermion over a full circle and back to its original
position, from which it acquires a minus sign [4], i.e.,
φ ¼ 0 → 6π∶ γ i → −γ i :
ð13Þ
Only for n ¼ 6 does each Majorana fermion come back to
its original position unchanged.
We now add up the contributions of various phase slips
to derive an effective Hamiltonian for Majorana fermions as
a function of the offset charge ng for each state of the
harmonic oscillator:
041038-5
SAGAR VIJAY, TIMOTHY H. HSIEH, AND LIANG FU
PHYS. REV. X 5, 041038 (2015)
FIG. 4. Schematic of a 2π phase slip on the central superconducting island in a hexagonal superconducting array on a TI surface, with
the phase of the central island indicated in each panel. When the phase difference between neighboring islands is π, the pair of Majorana
fermions on the shared edges couple [4] as indicated. The 2π phase slip permutes the Majorana fermions as shown, leading to the
transformation in Eq. (12).
H α ðng Þ ¼ ϵ0α þ
6
X
ðtα;n Ûn ei2πnng þ H:c:Þ:
ð14Þ
n¼1
Here, ϵ0α is the quantized energy of the harmonic oscillator
given by Eq. (11), which is the same for all internal states of
the Majorana fermions. The second term describes quantum phase slips: tα;n denotes the amplitude of the αth
energy level of the harmonic oscillator tunneling between
two potential minima that differ by 2πn, while Ûn is the
unitary operator acting on the Majorana fermions due to a
2πn phase slip. The coupling tα;n depends on the energy
barrier in the phase-slip event and canpbe
modulated by
ffiffiffiffiffiffiffiffiffiffiffi
tuning Ec =EJ ; for example, tα;1 ∝ e− 8EJ =Ec [44]. The
offset charge ng provides an Aharonov-Bohm flux proportional to the winding number n.
The Hamiltonian (14) is analogous to the Bloch
Hamiltonian that describes the band structure of a particle
hopping in a one-dimensional periodic potential, with the
offset charge ng playing the role of crystal momentum.
Importantly, the phase particle carries internal degrees of
freedom resulting from Majorana fermions γ 1 ; …; γ 6 that
are unique to our system. A phase slip that moves the phase
particle to a different potential minimum also permutes the
Majorana fermions as shown in Eqs. (12) and (13), similar
to a spinful particle hopping in the presence of a nonAbelian gauge field. These permutations are represented by
the unitary operators Ûn in the effective Hamiltonian (14)
acting on Majorana fermions. For example, the operator Û 1
that generates the transformation (12) is given by
Û 1 ¼
1 þ γ2γ3 1 þ γ4 γ5 1 − γ6 γ1
pffiffiffi
pffiffiffi
pffiffiffi
2
2
2
1 þ γ1γ2 1 þ γ3 γ4 1 þ γ5γ6
pffiffiffi
pffiffiffi :
× pffiffiffi
2
2
2
ð15Þ
It follows from the addition of phase slips that Û n ¼ ðÛ1 Þn .
In particular, the unitary operator Û3, which takes γ i to −γ i
as shown in Eq. (13), has a simple form
Û3 ¼ −
6
Y
γ i ¼ iO;
ð16Þ
i¼1
where O is the plaquette operator defined in the Majorana
plaquette model (2). On the other hand, for n ¼ 1, 2, 4, or 5,
Un is a sum of operators γ i γ j , γ i γ j γ k γ l , and iO.
Substituting the expressions for the U n ’s into Eq. (14),
we find that the effective Hamiltonian induced by the small
charging energy of a single island takes the following form:
Hα ðng Þ ¼ ϵ0α þ Δα ðng ÞO þ V α ðng Þ;
ð17Þ
with
Δα ðng Þ ¼
5
X
tα;m sinð2πmng Þ:
ð18Þ
m¼1
V α ðng Þ includes a constant tα;6 cosð12πng Þ, as well as
Majorana bilinear and quartic operators generated by phase
slips with winding number n ≠ 0 mod 3. Unlike O, these
operators on neighboring islands do not commute.
However, by appropriately tuning ng , the contribution of
quartic operators to the effective Hamiltonian may vanish,
so the only remaining terms in the Hamiltonian will be the
six-Majorana interaction and Majorana bilinear terms. The
bilinear term receives no contribution from any 6πm
phase slip, while Δα ðng Þ receives contributions from every
2πm phase-slip process. Therefore, for the remainder of
this work, we assume that V can be treated as a perturbation
to the Majorana plaquette model that does not destroy the
Z2 topological order of the gapped phase. We note that an
041038-6
MAJORANA FERMION SURFACE CODE FOR UNIVERSAL …
alternative setup without the presence of V has been
presented in a recent work [45].
C. Discussion
In deriving the effective Hamiltonian (14), we have
implicitly assumed that Majorana fermions are the only
low-lying excitations involved in phase-slip events, separated by an energy gap from other Andreev bound states in
the junctions between islands. This assumption is valid
because of the finite size of the islands, which leads to a
discrete Andreev bound-state spectrum with a finite gap for
all values of the phase. The presence of this gap justifies our
derivation of the effective Hamiltonian (14) in a controlled
manner.
Over the last few years, considerable experimental
progress has been made in hybrid TI-superconducting
systems. Proximity-induced superconductivity and supercurrents have been observed in a number of TI materials
[46–52]. Low-temperature scanning tunneling microscopy
(STM) experiments have found a proximity-induced superconducting gap on TI surface states, and the tunneling
spectrum of Abrikosov vortices shows a zero-bias conductance peak, which is robust in a range of a magnetic
field and splits at higher fields [12]. This peak has been
attributed to the predicted Majorana fermion zero modes in
the vortex cores of superconducting TI surface states. In
view of these rapid, unabated advances, we regard the
hybrid TI-superconductor system as a very promising
material platform for realizing the Majorana plaquette
model and studying the exciting physics of Majorana
fermions enabled by quantum phase slips.
Besides TIs, a two-dimensional electron gas (2DEG)
with spin-orbit coupling (such as InAs) can be driven into a
helical state with an odd number of spin-polarized Fermi
surfaces by an external Zeeman field, which provides
another promising platform for realizing topological superconductivity via the proximity effect [53,54]. In this
topological regime, vortices and trijunctions of a superconducting 2DEG host a single Majorana fermion, similar
to the TI surface. Thus, our proposed setup for the
Majorana plaquette model in Sec. III A directly applies
to this system as well.
In addition to TIs and 2DEG, (quasi-)one-dimensional
semiconductors and metals with strong spin-orbit coupling
have become a hotly pursued system to search for Majorana
fermions [5–7]. Signatures of Majorana fermions were
reported in 2012, based on the observation of a zero-bias
conductance peak in hybrid nanowire-superconductor systems [9,10]. One can envision a network of nanowires in
proximity to Cooper-pair boxes to realize our Majorana
plaquette model. In this direction, it is worth noting that a
new physical system—a nanowire with an epitaxially
grown superconductor layer—has recently been introduced
to study Andreev bound states in the presence of charging
energy [55].
PHYS. REV. X 5, 041038 (2015)
Many other physical systems for Majorana fermions
have been theoretically proposed and experimentally pursued; they are too numerous to list. Regardless of the
particular system, nonlocal interactions between multiple
Majorana fermions emerge from the charging energy of
superconductors via quantum phase slips, and in the
universal regime, such interactions are determined by the
transformation of Majorana fermions under phase slips, as
we have shown in Sec. III B.
Finally, we note several previous works related to our
Majorana plaquette model and its physical realization. In
Ref. [35], Xu and Fu first introduced a model of interacting
Majorana fermions that realizes Z2 topological order. This
model involves four-body and eight-body plaquette interactions on square and octagonal plaquettes in a twodimensional lattice. Physical realizations of this model
were proposed using an array of superconductor islands in
proximity to either a 2D TI [35] or semiconductor nanowires [56,57]. The four-body nonlocal interaction between
Majorana fermions comes directly from the charging
energy, whereas the eight-body interaction comes from a
high-order ring-exchange process generated by singleelectron tunneling between islands. In comparison, our
Majorana plaquette model on the honeycomb lattice has the
theoretical novelty of possessing an exact anyon permutation symmetry and can be realized in a much simpler
manner using an array of superconductors on a 3D TI with
global phase coherence, with all the required interactions
coming directly from the charging energy. We also note a
recent work on lattice models of Majorana fermions in
Abrikosov vortices on a superconducting TI surface [58],
which use different interactions and do not exhibit
topological order.
IV. MAJORANA SURFACE CODE
In the rest of this work, we demonstrate that the
Majorana plaquette model finds a natural application as
a Majorana fermion surface code, on which universal
quantum computation and error correction may be performed. The main idea of the surface code is to (i) use
anyons of the Majorana plaquette model to encode logical
qubits, (ii) manipulate anyons to perform gate operations
on logical qubits, and (iii) use commuting measurements of
the Majorana plaquette operators for error correction. We
describe the detailed implementation of the Majorana
surface code, including the creation of logical qubits, error
correction, and protocols for logical gate operations
required for universal quantum computation.
The surface code architecture [24,25,28] is a measurement-based scheme for quantum computation. It uses
projective measurements of commuting operators—called
“stabilizers”—acting on a 2D array of physical qubits to
produce a highly entangled “code state” jψi. Logical qubits
are created by stopping the measurement of certain commuting operators to create “holes.” The different possible
041038-7
SAGAR VIJAY, TIMOTHY H. HSIEH, AND LIANG FU
anyon charges at a hole are the degrees of freedom that
define a logical qubit. Logical gates are realized by
manipulating and braiding holes via a sequence of
measurements.
A key advantage of the surface code is its remarkable
capability for error detection. The random measurement of
an operator in the surface code corresponds to nucleating
pairs of anyons, a process that can be reliably measured by
tracking the eigenvalues of the commuting stabilizers.
Reliable error detection hinges on having (i) a large number
of physical qubits for a given encoded logical qubit and
(ii) a sufficiently low error rate for stabilizer measurements
[28]. For the previously studied surface code with bosonic
physical qubits, it has been estimated [59,60] that below a
threshold as high as 1% error rate per physical qubit
operation, scaling the size of the surface code permits an
exponential suppression of errors propagated. This error
tolerance makes the surface code architecture one of the
most realistic approaches to practical, large-scale quantum
computation.
Recent practical realizations of the surface code have
used superconducting qubits coupled to a microwave
transmission line resonator to perform qubit manipulations
and measurements [31–33]. Here, a physical qubit is
defined by two energy levels arising from quantization
of phase fluctuations in a conventional Cooper-pair box.
The surface code is implemented on a 2D array of physical
qubits with the four-qubit interactions of Kitaev’s toric code
Hamiltonian [22] as the set of commuting stabilizers. The
four-qubit stabilizer is measured by performing a sequence
of single- and two-qubit gates between the four physical
qubits and additional ancilla qubits [28]. Experiments have
demonstrated the remarkable ability to operate these
physical gates with fidelity above the threshold required
for surface code error correction [31]. Recent experiments
have also used error detection to preserve entangled code
states on a surface code with a 9 × 1 [32] and a 2 × 2 [33]
array of stabilizers. It remains to be shown that logical
qubits can be successfully encoded and manipulated via
logical gates in these surface code arrays.
A. Implementation
We implement the Majorana surface code on a 2D array
of Majorana fermions by performing projective measurements of the Majorana plaquette operators fOp g, which
form a complete set of commuting stabilizers. For the
remainder of this paper, we will use “plaquette operators”
and “stabilizers” interchangeably to refer to fOp g. A
practical physical system for implementing the Majorana
surface code is the superconductor-TI hybrid system
introduced in the previous section. We place an array of
superconducting islands on the TI surface, which are
strongly Josephson coupled. By introducing external circuits or applying fluxes, we engineer the Josephson
coupling between islands to achieve the phase
PHYS. REV. X 5, 041038 (2015)
FIG. 5. (a) Schematic of the harmonic oscillator energy levels of
the effective Hamiltonian (17), centered at φ ¼ 2πn, with the 2π
and 4π phase-slip amplitudes for the lowest energy levels shown.
In (b), we show a schematic plot of the two lowest harmonic
oscillator levels as a function of the gate charge. The energy
splittings Δ1 and Δ2 are between states with even (Γ ¼ þ1) and
odd fermion parity (Γ ¼ −1) within the first and second harmonic
oscillator levels, respectively. Each level within a fixed fermion
parity sector is nearly four-fold degenerate.
configuration in Fig. 3, leading to a honeycomb lattice
of Majorana fermions at trijunctions.
To perform a projective measurement of the Majorana
plaquette operator on a given island, i.e., a single stabilizer,
we decrease the Josephson coupling of the island with the
rest of the array to activate quantum phase slips from the
small but nonzero charging energy on this island. As shown
by the effective Hamiltonian in Eq. (17), these quantum
phase slips (partially) lift the degeneracy between states in
the eight-dimensional Fock space of the six Majorana
fermions. In particular, for every energy level of the
harmonic oscillator, there is an energy splitting Δα ðng Þ
between states of Majorana fermions with Γ ¼ þ1 (even
fermion parity) and with Γ ¼ −1 (odd fermion parity) from
Eq. (18), where Γ is the stabilizer eigenvalue; this is shown
schematically in Fig. 5(b). Therefore, the charging energy
of the island creates an energy difference between different
stabilizer eigenstates. Furthermore, the energy gap between
the two lowest harmonic oscillator levels on the island is a
function of the stabilizer eigenvalue Γ ¼ 1, and in the
limit of negligible interaction V, it takes the following
form:
041038-8
MAJORANA FERMION SURFACE CODE FOR UNIVERSAL …
ΔEΓ ðng Þ ¼ ϵ0 þ ½Δ2 ðng Þ − Δ1 ðng ÞΓ þ ;
ð19Þ
pffiffiffiffiffiffiffiffiffiffiffiffiffi
where ϵ0 ≡ ϵ02 − ϵ01 ≈ 8EJ Ec . The sensitivity of the
energy gap to the stabilizer eigenvalue now permits a
stabilizer measurement by simply measuring the energy
gap. By shining a probe microwave beam on this island, we
may measure the phase shift of the transmitted photons to
determine the gap between the two harmonic oscillator
levels [61,62].
We now perform these stabilizer measurements on all of
the superconducting islands to project onto an eigenstate of
the Majorana plaquette Hamiltonian (1); this will be our
reference “code state.” We continue to perform measurements on all hexagonal islands in each cycle of the surface
code in order to maintain the state. In subsequent cycles, we
may encode logical qubits into the code state and manipulate the qubits via measurement. While projection onto the
code state and error correction in the surface code rely
exclusively on measuring the six-Majorana plaquette interaction, manipulation of logical qubits also requires measuring nearest-neighbor Majorana bilinears on the
hexagonal lattice. This may be done by tuning the phase
of neighboring superconducting islands to bring the pair of
Majorana fermions on the shared edge sufficiently close
together [4] so that the resulting wave-function overlap
further splits the nearly four-fold degeneracy within a
single fermion parity sector, shown in Fig. 5(b). Again,
the Majorana bilinear may be measured by shining a probe
beam to measure the energy gap to the next harmonic
oscillator level.
Using the commuting six-Majorana operators in our
plaquette model to realize a surface code provides unique
advantages over the more conventional surface code with
bosonic physical qubits. First, while a four-spin stabilizer
measurement in the usual surface code requires performing
6–8 gates or measurements between a set of physical and
ancilla qubits [28,59], stabilizer eigenvalues in the
Majorana surface code are obtained via a single-step
measurement by shining a probe beam. We emphasize
that even when measurement is not being performed, the
intrinsic charging energy of the islands generates a finite
gap Δ1 ðng Þ to creating anyon excitations and naturally
suppresses errors at temperatures kB T < Δ1 ðng Þ. We anticipate that the corresponding error tolerance for scalable
quantum computation is substantially improved for the
Majorana surface code. Second, the Majorana surface code
operates with lower overhead than its bosonic counterpart,
using three-qubit stabilizers and requiring no ancilla qubits.
Finally, the anyon transmutation required to perform a
logical Hadamard gate in the conventional surface code
corresponds to a duality transformation that exchanges the
star and plaquette toric code operators. This operation is
quite difficult to perform on a single logical qubit as it also
requires lattice surgery to patch the transformed logical
qubit back into the remaining surface code [28,63].
PHYS. REV. X 5, 041038 (2015)
As lattice symmetries permute anyon sectors in the
Majorana plaquette model, anyon transmutation in the
Majorana surface code corresponds to a lattice translation
of the logical qubit, substantially simplifying the Hadamard
gate implementation.
B. Logical qubits and error correction
Logical qubits may be encoded in the surface code by
ceasing the measurement of the plaquette operator on a
hexagonal superconducting island in a surface code cycle,
while continuing measurements on all other plaquettes. In
theory, we could stop measuring a single plaquette and
define a two-level system, with the Ẑ and X̂ operators of the
logical qubit defined by the plaquette operator and a Wilson
line connecting the plaquette to the boundary, respectively.
A pair of such qubits on the A-type plaquettes is shown in
Fig. 6(a), where the solid and dashed lines correspond to
products of Majorana fermions that define the indicated
logical operators. The two qubits shown may also be
coherently manipulated by acting with the Wilson line
operator connecting the two plaquettes, denoted X̂12 .
In practice, however, it is difficult to manipulate qubits
with an operator that connects to a distant boundary, so it is
simpler to encode a logical qubit by stopping the stabilizer
measurement on two well-separated plaquettes of the same
type. We choose to only manipulate 2 of the 4 resulting
degrees of freedom by defining Ẑ ≡ Op and X̂ ≡ W pq , the
Wilson line operator connecting the two plaquettes. We use
the opposite convention to define the logical Ẑ and X̂
operators for a qubit on the adjacent B plaquettes; an
example of such logical qubits is shown in Fig. 6(b). We
note that when such a qubit is created, it is automatically
initialized to an eigenstate of the plaquette operator, with
the eigenvalue given by the measurement performed in the
previous surface code cycle. As a result, logical qubits of
type A (B) are initialized to an eigenstate of the Ẑ (X̂)
logical operator.
To reduce errors during qubit manipulation, we may
define a qubit by ceasing measurement of multiple adjacent
plaquettes as shown in Fig. 6(b). In this particular case, the
logical operator X̂ is still a Wilson line connecting to
another set of distant holes. However, the logical Z is
defined as Ẑ ≡ Op ⊗ Oq ⊗ Or . For the remainder of our
discussion, we consider logical qubits with only a single
plaquette operator used to define the logical Ẑ; the
generalization to larger qubits is straightforward.
Errors may occur during qubit manipulation, including
(1) single-qubit errors due to the unintended measurement
of a local operator involving an even number of Majorana
fermions and (2) measurement errors. Single-qubit error
correction may be performed on logical qubits by constantly measuring the remaining plaquette eigenvalues
during surface code cycles. Since only pairs of plaquettes
may be flipped simultaneously by a random measurement,
corresponding to the nucleation of a pair of anyons of a
041038-9
SAGAR VIJAY, TIMOTHY H. HSIEH, AND LIANG FU
PHYS. REV. X 5, 041038 (2015)
registering the eigenvalue of a plaquette operator; these
are naturally corrected by performing multiple surface code
cycles to verify the accuracy of a measurement.
C. Logical gate implementations
The Majorana surface code may be used for universal
quantum computation by implementing CNOT, T, and
Hadamard gates on logical qubits; this has been extensively
studied in the context of the surface code architecture with
underlying bosonic degrees of freedom [28,63]. Here, we
describe the implementations of these gates in our realization of quantum computation with a Majorana surface
code. Our gate implementations follow the spirit of the
implementations presented in Ref. [28].
All gates in the Majorana surface code are implemented
on logical qubits via a sequence of measurements. Let Û be
the desired unitary we wish to perform on the quantum
state of several logical qubits, defined by the logical
operators fX̂i g and fẐi g. It is convenient to keep track
of the transformation of the logical state by monitoring
the transformation of logical operators X̂i → ÛX̂i Û† ,
Ẑi → Û Ẑi Û† . In practice, performing the appropriate
sequence of measurements will yield the transformation
W, such that
FIG. 6. Logical qubits in the Majorana surface code. In (a) we
stop the measurement of two plaquette operators in subsequent
surface code cycles, increasing the ground-state degeneracy by a
factor of 4. If we take Ẑ1 and Ẑ2 to be the logical Ẑ operators for
the two encoded qubits, the corresponding X̂ 1 and X̂ 2 operators
are given by Wilson lines connecting to the boundary. The two
qubits may be coherently manipulated by applying the operator
X̂ 12 as shown. In practice, it is simpler to define logical qubits by
stopping the measurement of pairs of plaquettes of a single type,
with the logical X̂ and Ẑ defined as shown in (b). We may also
consider a logical qubit made of several “holes,” as in (c), to
minimize errors during qubit manipulation.
single type, detecting the change of an odd number of
plaquette eigenvalues in a single surface code cycle will
generally signal the presence of a random measurement
performed on a nearby logical qubit. More precisely, when
a stabilizer eigenvalue changes in a surface code cycle, it is
efficient to store the location of that stabilizer and wait
several code cycles, accumulating a spacetime diagram of
stabilizer errors as additional errors occur [30,59,60]. After
sufficiently many code cycles, the spacetime diagram may
be used to determine the most likely configuration of
Wilson lines that could have generated those errors [28,29]
using a minimum-weight perfect matching algorithm
[64,65]. Errors may be subsequently corrected by software
when performing logical qubit manipulations and readouts
[28]. Random measurement errors involve incorrectly
ÛX̂i Û† ¼ Ŵ X̂i Ŵ † ;
ð20Þ
Û Ẑi Û† ¼ Ŵ Ẑi Ŵ † ;
ð21Þ
where the signs depend on the outcomes of the specific
measurements performed. These measurement outcomes
are stored in software and used to correctly interpret the
readout of a logical qubit.
In what follows, we often demonstrate our gate implementations in an “operator picture,” where a set of
operators in the surface code ô1 ; …; ôn and p̂1 ; …; p̂m
with eigenvalues 1 are measured in an appropriate
sequence. This implements a logical gate via the desired
transformations:
Ẑ → Ẑ ⊗
n
Y
ôj ¼ Û Ẑ Û † ;
ð22Þ
p̂j ¼ Û X̂ Û † :
ð23Þ
j¼1
X̂ → X̂ ⊗
m
Y
j¼1
In practice, the measured outcomes for the fôi g and fp̂j g
operators will be stored by software and used to obtain the
above transformations during logical qubit readout.
CNOT gate.—A CNOT gate takes two qubits, a
“control” and a “target,” and flips the value of the target
qubit based on the value of the control; it then returns the
041038-10
MAJORANA FERMION SURFACE CODE FOR UNIVERSAL …
PHYS. REV. X 5, 041038 (2015)
control unchanged. The action of a CNOT takes the
following form in the basis of two-qubit states:
0
1
0 0
B0
B
Ĉ ¼ B
@0
1
0 0
0C
C
C:
1A
0 1
0
1 0
0
0
ð24Þ
A CNOT gate may be implemented by braiding logical
qubits in the Majorana surface code. In the simplest case, a
CNOT between two logical qubits of different types is
implemented through a single braiding operation that
produces an overall sign if the hexagonal ends of both
qubits contain an anyon, due to the π mutual statistics,
demonstrated in Sec. II. In the following section, we first
describe the procedure to move a logical qubit along a
given type of plaquette before discussing the braiding
procedure required to produce a CNOT gate.
Consider the A-type logical qubit shown in Fig. 7. To
move the qubit one unit to the right, we perform the
following sequence of measurements. We begin by multiplying the Ẑ logical operator by the eigenvalue of the
adjacent r plaquette operator to perform the transformation
0
Ẑ → Ẑ ≡ Ẑ ⊗ Ôr :
ð25Þ
As the r plaquette is being continuously measured, its
eigenvalue Or ¼ 1 is known from the previous surface
code cycle. In the next cycle, we stop measuring Ôr and
measure the Majorana bilinear iγη. We then multiply the X̂
operator by the measurement outcome, affecting the
transformation
X̂ → X̂0 ≡ X̂ ⊗ iγη:
ð26Þ
In the final surface code cycle, we begin measuring the
original Ẑ stabilizer and continue to include the measurement of the Ẑ stabilizer in all subsequent surface code
cycles. Furthermore, we redefine the logical operator Ẑ0 as
Ẑ0 → Ẑ00 ≡ Ẑ0 ⊗ Ẑ:
ð27Þ
The initial qubit configuration and final outcome are
depicted schematically in Fig. 7. This sequence of measurements has shifted the A-type qubit by moving its
hexagonal end one unit to the right, and it may generally
be used to move an A- or a B-type logical qubit within the A
or B plaquettes, respectively.
We may now braid pairs of logical qubits to perform a
CNOT gate in the Majorana surface code. The simplest
CNOT that we may realize is between two distinct types of
qubits, taking the A qubit as the control, as shown in Fig. 8.
Since the qubits are distinct, braiding the B-type qubit—
with logical operators X̂B and ẐB —along a closed path l,
FIG. 7. We may move a logical qubit defined by X̂ and Ẑ
operators along a given sublattice. We first multiply the logical Ẑ
by Ôr. After measuring iγη in the next code cycle, we extend the
logical X̂ → X̂ ⊗ iγη. Finally, we begin measuring Ẑ in the next
surface code cycle and restore Ôp and Ôq to six-Majorana
operators.
enclosing the second qubit, (i) multiplies the Wilson line of
the B-type qubit by the anyon charge enclosed by l and
(ii) multiplies the Wilson line of the A qubit by the anyon
charge of the B qubit. This results in the transformation
X̂A → X̂A ⊗ X̂B ;
ẐB → ẐA ⊗ ẐB ⊗
Y
Ôp ;
ð28Þ
p∈l
where fÔp g are A- and C-type plaquette operators enclosed
by the braiding trajectory, as shown in Fig. 8. Since the
eigenvalues of the enclosed plaquette operators are known
from the previous surface code cycle, we may implement
the logical CNOT (ẐA → ẐA , ẐB → ẐA ⊗ ẐB ) by multiplying the transformed ẐB by an appropriate sign. In
summary, the simplest braiding process between an A
and a B logical qubit implements a CNOT on the B qubit,
with the A qubit as the control.
A CNOT between two logical qubits of the same type
may also be performed by appropriately braiding pairs of
distinct types of logical qubits. In this case, we take one
qubit as the control by convention and store the outcome of
the CNOT gate in a third ancilla qubit. First, consider
performing a CNOT gate on two A-type qubits. To implement the CNOT, we prepare two additional ancilla qubits;
the first is an
pffiffiffiA qubit prepared in the state jφi ≡
½jþz i þ j−z i= 2, while the second is a B qubit prepared
in the state jþx i, with jz i and jx i the eigenstates of the
logical Z and X operators, respectively. Both ancilla qubits
are prepared by measuring a þ1 eigenvalue for the Wilson
line joining the pair of plaquettes of the appropriate qubit.
For the A (B) qubit, this projects onto an eigenstate of the
041038-11
SAGAR VIJAY, TIMOTHY H. HSIEH, AND LIANG FU
FIG. 8. CNOT gate. Braiding two logical qubits to perform a
logical CNOT. In (a), a possible trajectory for braiding the first
qubit around the second is indicated by the dotted line. Since
the two qubits live on distinct sublattices, the braiding procedure
induces the transformation X̂ A → X̂ A ⊗ X̂ B and ẐB →
ẐA ⊗ ẐB ⊗ Ô, where Ô is the product of the colored plaquettes
shown. As a result, this operation performs a CNOT transformation on the braided qubit.
logical X (Z) operator and produces the desired ancilla
states.
We now represent a complete basis of the four-qubit
states as jzB ; zc ; zt ; zout i, referring to the eigenvalues of the
logical Z operators of the B ancilla, the control, the target,
and the ancilla A qubits, respectively. We start out with an
initial state jψ init i of the form
1
jψ init i ≡ pffiffiffi ½jþ; zc ; zt ; þi þ jþ; zc ; zt ; −i:
2
ð29Þ
Next, we braid the B ancilla qubit around all three
remaining qubits as shown in Fig. 9(a). Up to an overall
sign determined by the eigenvalues of plaquette operators
enclosed by the braiding trajectory that are known from
previous surface code cycles, this braiding implements the
transformation ẐB → ẐB ⊗ Ẑc ⊗ Ẑt ⊗ Ẑout on the logical
PHYS. REV. X 5, 041038 (2015)
FIG. 9. Braiding processes that implement the transformation
(a) Ẑa → Ẑa ⊗ Ẑc ⊗ Ẑt ⊗ Ẑout up to an overall sign, as determined by the product of the appropriate plaquette operators
enclosed by the path l, and (b) Ẑout → Ẑout ⊗ ẐA , Ẑt → Ẑt ⊗ ẐA ,
Ẑc → Ẑc ⊗ ẐA . The two braids are used to realize CNOT gates
between two (a) A-type and (b) B-type logical qubits, respectively. By convention, we take the lowest qubit enclosed by the
braiding trajectory to be the control for the logical CNOT.
Z of the ancilla qubit, where Ẑc , Ẑt , and Ẑout are the logical
Z operators for the control, target, and ancilla A-type
qubits, respectively. The final state we obtain is then of
the form
1
jψ final i ¼ pffiffiffi ½jzc zt ; zc ; zt ; þi þ j−zc zt ; zc ; zt ; −i:
2
ð30Þ
This braiding process is convenient, as a measurement of
the state of the B qubit can determine whether the state of
the A ancilla contains the correct outcome of the CNOT
operation. If we now measure the logical Z of the B qubit
and obtain ẐB ¼ þ1, then we project onto a state with
zc zt ¼ zout . In this case, the A ancilla qubit contains the
correct outcome of the CNOT between the other A qubits. If
ẐB ¼ −1, however, then zc zt ¼ −zout and the A ancilla
contains the opposite of the correct CNOT outcome. In this
case, we may act with X̂out on the A ancilla qubit in the
surface code software [28] to obtain the desired final state.
041038-12
MAJORANA FERMION SURFACE CODE FOR UNIVERSAL …
PHYS. REV. X 5, 041038 (2015)
A similar process may be used to perform logical
CNOT’s between two B qubits; now, we prepare an A
ancilla qubit and a B ancilla qubit in the states shown in
Fig. 9(b). After braiding the ancilla A qubit around the
control, target, and ancilla B qubits, if we measure
X̂A ¼ þ1, then the B ancilla contains the desired outcome
of the CNOT operation. Again, by convention, we take the
control qubit to be the first one enclosed by the braiding
trajectory, as shown in Fig. 9(b).
Hadamard gate.—The Hadamard gate is a single-qubit
gate taking the matrix form
1
Ĥ ¼ pffiffiffi
2
1
1
1 −1
:
ð31Þ
The action of a Hadamard gate is to exchange the logical X̂
and Ẑ operators so that Ĥ X̂ Ĥ† ¼ Ẑ and Ĥ Ẑ Ĥ† ¼ X̂. As
the logical X̂ and Ẑ are defined oppositely on different
types of qubits, a Hadamard operation in the bosonic
surface code corresponds to an electric-magnetic duality
transformation that interchanges star and plaquette operators in the toric code. In the ordinary surface code, such a
transformation is quite difficult to implement, requiring a
series of Hadamards on physical qubits enclosing the
logical qubit so as to interchange the X̂ and Ẑ stabilizers,
followed by physical swap gates in order to correctly patch
the transformed logical qubit back into the remaining
surface code array [63]. As lattice symmetries permute
the anyons in the Majorana plaquette model, however, the
logical Hadamard gate may be realized in the Majorana
surface code by simply moving a logical qubit between
distinct plaquettes.
We implement the logical Hadamard by the procedure
shown in Fig. 10. Consider an A-type logical qubit. We
multiply the logical X̂ operator of the qubit, defined by the
Wilson line in Fig. 10(a), by the product of the adjacent
plaquette operators fÔμk g extending between the hexagonal ends of the qubit. The eigenvalues of these plaquette
operators are known from previous surface code cycles.
This operation implements the transformation
Y
X̂ → Ẑ0 ≡ X̂ ⊗
ð32Þ
Ôμk :
k
At the same time, we multiply the logical Ẑ by the adjacent
plaquette operator Ôp, shown in Fig. 10(a), that borders the
logical qubit above:
Ẑ → X̂0 ≡ Ẑ ⊗ Ôp :
ð33Þ
In subsequent surface code cycles, we stop measuring the
eigenvalue of Ôp . We implement a similar transformation
on the other hexagonal end of the logical qubit by stopping
the measurement of the plaquette operator above the
other qubit hole. The end result, after performing these
FIG. 10. Hadamard gate. A logical Hadamard is performed by
transferring a qubit between distinct sublattices so that the logical
X̂ and Ẑ operators are exchanged. We do this by taking the qubit
in (a) and multiplying the logical X̂ by the plaquette operators
fÔμk g and the logical Ẑ by Ôp and ceasing measurement of the
fermion parity of plaquette p, yielding the operators shown in (b).
Next, we measure the product ðiη1 η2 Þðiη3 η4 Þ… and Ôq and
multiply by Ẑ0 and X̂ 0 , respectively. The final result is shown
in (d).
operations, is shown in Fig. 10(b). The solid and dashed
blue lines indicate the products of the Majorana fermions
on the appropriate sites that define the X̂0 and Ẑ0 operators,
respectively.
In the next surface code cycle, we measure the product
ðiη1 η2 Þðiη3 η4 Þ… of the Majorana fermions along the lower
“string” that defines the Ẑ0 operator; this measurement
commutes with X̂0 since the two operators do not overlap,
as shown in Fig. 10(c). Afterwards, we measure Ôq , as well
041038-13
SAGAR VIJAY, TIMOTHY H. HSIEH, AND LIANG FU
PHYS. REV. X 5, 041038 (2015)
as the plaquette operator Ôh for the other hole of the
original logical qubit. Then, we may perform the following
transformations on the logical X̂0 and Ẑ0 operators:
X̂0 → X̂00 ≡ X̂0 ⊗ Ôq ;
Ẑ0 → Ẑ00 ≡ Ẑ0 ⊗
ð34Þ
Y
ðiη2l−1 η2l Þ:
ð35Þ
l
This yields the logical qubit shown in Fig. 10(d). In
subsequent surface code cycles, we continue measuring
the eigenvalues of Ôq and Ôh . Since the logical Ẑ and X̂
operators are defined differently on the A and B-type
plaquettes, our procedure for transforming our A qubit
into a B qubit implements a logical Hadamard gate. An
identical protocol may be used to perform a Hadamard gate
on a logical B qubit.
S and T gates.—Finally, we implement the logical S and
T gates, described by the following single-qubit operations:
Ŝ ¼
1 0
0
i
;
T̂ ¼
1
0
:
eiπ=4
0
Depending on the measurement outcome, we implement an
S gate as shown below:
ð36Þ
As demonstrated in Ref. [28], it is possible to realize these
gates by performing a series of logical Hadamard and
CNOT gates between the logical qubit and an appropriate
logical ancilla qubit. Here, we first discuss the S- and
T-gate implementations, given the appropriate ancilla
qubit, before outlining a procedure for creating these
logical ancillas in the surface code.
To implement an S gate, we prepare a logical ancilla in
the state
1
jφS i ≡ pffiffiffi ½jþz i þ ij−z i:
2
FIG. 11. S- and T-gate ancilla preparation. We create the jφS i
and jφT i ancilla states, which are needed to realize logical S and
T gates by preparing the short qubit [28] shown above. We cease
stabilizer measurements on two adjacent plaquettes p and q. In
the next surface code cycle, we perform a rotation of the two-level
system defined by iγη. Finally, we enlarge the logical qubit by
extending one end of the qubit to guarantee stability against
noise.
ð37Þ
Then, if jΨi is the state of the logical qubit of interest, the
following sequence of logical Hadamard and CNOT gates
implements the transformation jΨi → ŜjΨi [28]:
If the measurement outcome MZ ¼ þ1, then we obtain
the correct output T̂jΨi; otherwise, if MZ ¼ −1, then we
have performed the transformation jΨi → X̂ T̂ † jΨi. In this
case, we implement an S gate on the logical qubit and
obtain the final state iX̂ Ẑ T̂ jΨi. The action of the operator
iX̂ Ẑ may be undone in the surface code software to
implement a pure T gate on the logical qubit [28].
To realize the above implementations, we may prepare
logical ancilla qubits in the states jφT i and jφS i as follows.
First, we create a “short qubit” [28] by ceasing the fermion
parity measurement on two adjacent plaquettes p, q
belonging to the same sublattice, as shown in Fig. 11(a).
For this qubit, let X̂ ≡ Ôp and Ẑ ≡ iγη. The qubit is
initialized to a state jΨ i such that X̂jΨ i ¼ jΨi. In a
basis of eigenstates of the logical
pffiffiffi Ẑ, the qubit state takes the
form jΨ i ¼ ðjþz i j−z iÞ= 2. Now, we assume that the
two-level system formed by the pair of Majorana fermions
γ and η can be manipulated by performing a rotation
R̂ðθÞ ¼
To perform a T gate, we first prepare a logical ancilla in
the state
1
jφT i ≡ pffiffiffi ½jþz i þ eiπ=4 j−z i:
2
ð38Þ
The T gate is then implemented via a probabilistic circuit.
We perform a CNOT between the ancilla and the logical
qubit of interest and then measure the logical Ẑ of the qubit.
1
0
0 eiθ
ð39Þ
that acts in the basis of jz i states. This may be implemented by using the phase of the adjacent superconducting
islands to tune the coupling between the Majorana zero
modes [66]. To prepare the state jφS i, we perform the
rotation R̂ðπ=2ÞjΨ i in the next surface code cycle,
while to prepare jφT i, we perform the operation
041038-14
MAJORANA FERMION SURFACE CODE FOR UNIVERSAL …
R̂ðð2π πÞ=4ÞjΨ i. Afterwards, to guarantee the stability
of the qubit against noise generated by the environment, we
increase the length of the logical X̂ operator by extending
one end of the logical qubit, as discussed in detail
previously and shown schematically in Fig. 7. In practice,
a high-fidelity implementation of the S and T gates requires
that the short qubits are put through a distillation circuit, as
discussed in Ref. [28], which may be implemented using a
sequence of logical CNOT gates with other ancilla logical
qubits.
We have presented a two-dimensional model of interacting Majorana fermions that realizes a novel type of Z2
topological order with a microscopic S3 anyon symmetry.
The required multifermion interactions in the plaquette
model are naturally generated by phase slips in an array of
phase-locked s-wave superconducting islands on a TI
surface. Based on this physical realization, we propose
the Majorana surface code and provide the necessary
measurement protocols and gate implementations for
universal quantum computation. The Majorana surface
code provides substantially increased error tolerance,
reduced overhead, and simpler logical gate implementations over a surface code with bosonic physical qubits. We
are optimistic that the Majorana fermion surface code will
be physically implemented and may provide an advantageous platform for fault-tolerant quantum computation.
ACKNOWLEDGMENTS
We thank Patrick Lee for helpful comments and discussion. This work was supported by the Packard
Foundation (L. F.), and the DOE Office of Basic Energy
Sciences, Division of Materials Sciences and Engineering
under Grant No. DE-SC0010526 (S. V. and T. H.). S. V. and
L. F. conceived and developed the Majorana fermion
surface code for universal quantum computation and wrote
the manuscript. T. H. contributed to the analysis of the
Majorana plaquette model.
[1] E. Majorana, Teoria Simmetrica dell’elettrone e del
Positrone, Nuovo Cimento, 14, 171 (1937).
[2] J. Alicea, New Directions in the Pursuit of Majorana
Fermions in Solid State Systems, Rep. Prog. Phys. 75,
076501 (2012).
[3] C. W. J. Beenakker, Search for Majorana Fermions in
Superconductors, Annu. Rev. Condens. Matter Phys. 4,
113 (2013).
[4] L. Fu and C. L. Kane, Superconducting Proximity Effect and
Majorana Fermions at the Surface of a Topological
Insulator, Phys. Rev. Lett. 100, 096407 (2008).
[5] R. M. Lutchyn, J. D. Sau, and S. Das Sarma, Majorana
Fermions and a Topological Phase Transition in
Semiconductor-Superconductor Heterostructures, Phys.
Rev. Lett. 105, 077001 (2010).
PHYS. REV. X 5, 041038 (2015)
[6] Y. Oreg, G. Refael, and F. von Oppen, Helical Liquids and
Majorana Bound States in Quantum Wires, Phys. Rev. Lett.
105, 177002 (2010).
[7] A. C. Potter and P. A. Lee, Multichannel Generalization of
Kitaev’s Majorana End States and a Practical Route to
Realize Them in Thin Films, Phys. Rev. Lett. 105, 227003
(2010).
[8] S. Nadj-Perge, I. K. Drozdov, B. A. Bernevig, and A.
Yazdani, Proposal for Realizing Majorana Fermions in
Chains of Magnetic Atoms on a Superconductor, Phys. Rev.
B 88, 020407(R) (2013).
[9] V. Mourik, K. Zuo, S. M. Frolov, S. R. Plissard, E. P. A. M.
Bakkers, and L. P. Kouwenhoven, Signatures of Majorana
Fermions in Hybrid Superconductor-Semiconductor
Nanowire Devices, Science 336, 1003 (2012).
[10] A. Das, Y. Ronen, Y. Most, Y. Oreg, M. Heiblum, and
H. Shtrikman, Zero-Bias Peaks and Splitting in an Al-InAs
Nanowire Topological Superconductor as a Signature of
Majorana Fermions, Nat. Phys. 8, 887 (2012).
[11] S. Nadj-Perge, I. K. Drozdov, J. Li, H. Chen, S. Jeon, J. Seo,
A. H. MacDonald, B. Andrei Bernevig, and A. Yazdani,
Observation of Majorana Fermions in Ferromagnetic Atomic
Chains on a Superconductor, Science 346, 602 (2014).
[12] J.-P. Xu, M.-X. Wang, Z. L. Liu, J.-F. Ge, X. Yang, C. Liu,
Z. A. Xu, D. Guan, C. L. Gao, D. Qian, Y. Liu, Q.-H. Wang,
F.-C. Zhang, Q.-K. Xue, and J.-F. Jia, Experimental
Detection of a Majorana Mode in the Core of a Magnetic
Vortex inside a Topological Insulator-Superconductor
Bi2 Te3 =NbSe2 Heterostructure, Phys. Rev. Lett. 114,
017001 (2015).
[13] K. T. Law, P. A. Lee, and T. K. Ng, Majorana Fermion
Induced Resonant Andreev Reflection, Phys. Rev. Lett. 103,
237001 (2009).
[14] J. Liu, A. C. Potter, K. T. Law, and P. A. Lee, Zero-Bias
Peaks in the Tunneling Conductance of Spin-Orbit-Coupled
Superconducting Wires with and without Majorana
End-States, Phys. Rev. Lett. 109, 267002 (2012).
[15] A. Kitaev, Unpaired Majorana Fermions in Quantum
Wires, Phys. Usp. 44, 131 (2001).
[16] H.-J. Kwon, K. Sengupta, and V. M. Yakovenko, Fractional
ac Josephson Effect in p- and d-wave Superconductors, Eur.
Phys. J. B 37, 349 (2003).
[17] L. Fu and C. L. Kane, Josephson Current and Noise
at a Superconductor-Quantum Spin Hall InsulatorSuperconductor Junction, Phys. Rev. B 79, 161408(R)
(2009).
[18] L. Fu, Electron Teleportation via Majorana Bound States in
a Mesoscopic Superconductor, Phys. Rev. Lett. 104,
056402 (2010).
[19] G. Moore and N. Read, Nonabelions in the Fractional
Quantum Hall Effect, Nucl. Phys. B360, 362 (1991).
[20] N. Read and D. Green, Paired States of Fermions in Two
Dimensions with Breaking of Parity and Time-Reversal
Symmetries and the Fractional Quantum Hall Effect, Phys.
Rev. B 61, 10267 (2000).
[21] D. A. Ivanov, Non-Abelian Statistics of Half-Quantum
Vortices in p-Wave Superconductors, Phys. Rev. Lett. 86,
268 (2001).
[22] A. Kitaev, Fault-Tolerant Quantum Computation by
Anyons, Ann. Phys. (Amsterdam) 303, 2 (2003).
041038-15
SAGAR VIJAY, TIMOTHY H. HSIEH, AND LIANG FU
[23] C. Nayak, S. H. Simon, A. Stern, M. Freedman, and
S. Das Sarma, Non-Abelian Anyons and Topological
Quantum Computation, Rev. Mod. Phys. 80, 1083
(2008).
[24] S. Bravyi and A. Kitaev, Quantum Codes on a Lattice with
Boundary, Quantum Computers and Computing 2, 43
(2001).
[25] M. H. Freedman and D. A. Meyer, Projective Plane and
Planar Quantum Codes, Found. Comput. Math. 1, 325
(2001).
[26] R. Raussendorf and J. Harrington, Fault-Tolerant Quantum
Computation with High Threshold in Two Dimensions,
Phys. Rev. Lett. 98, 190504 (2007).
[27] R. Raussendorf, J. Harrington, and K. Goyal, Topological
Fault-Tolerance in Cluster State Quantum Computation,
New J. Phys. 9, 199 (2007).
[28] A. G. Fowler, M. Mariantoni, J. M. Martinis, and A. N.
Cleland, Surface Codes: Towards Practical Large-Scale
Quantum Computation, Phys. Rev. A 86, 032324 (2012).
[29] A. G. Fowler, A. M. Stephens, and P. Groszkowski, HighThreshold Universal Quantum Computation on the Surface
Code, Phys. Rev. A 80, 052312 (2009).
[30] E. Dennis, A. Kitaev, A. Landahl, and J. Preskill,
Topological Quantum Memory, J. Math. Phys. (N.Y.) 43,
4452 (2002).
[31] R. Barends, J. Kelly, A. Megrant, A. Veitia, D. Sank, E.
Jeffrey, T. White, J. Mutus, A. Fowler, B. Campbell et al.,
Superconducting Quantum Circuits at the Surface Code
Threshold for Fault Tolerance, Nature (London) 508, 500
(2014).
[32] J. Kelly, R. Barends, A. Fowler, A. Megrant, E. Jeffrey, T.
White, D. Sank, J. Mutus, B. Campbell, Y. Chen et al.,
State Preservation by Repetitive Error Detection in a
Superconducting Quantum Circuit, Nature (London) 519,
66 (2015).
[33] A. D. Córcoles, E. Magesan, S. J. Srinivasan, A. W. Cross,
M. Steffen, J. M. Gambetta, and J. M. Chow, Demonstration
of a Quantum Error Detection Code Using a Square Lattice
of Four Superconducting Qubits, Nat. Commun. 6, 6979
(2015).
[34] S. Bravyi, B. Terhal, and B. Leemhuis, Majorana Fermion
Codes, New J. Phys. 12, 083039 (2010).
[35] C. Xu and L. Fu, Fractionalization in Josephson Junction
Arrays Hinged by Quantum Spin Hall Edges, Phys. Rev. B
81, 134435 (2010).
[36] See Supplemental Material at http://link.aps.org/
supplemental/10.1103/PhysRevX.5.041038 for a derivation
of the braiding and self-statistics of the topological excitations in the Majorana plaquette model and a discussion
of how lattice symmetries in the model provide a microscopic realization of the S3 anyon symmetry.
[37] X.-G. Wen, Quantum Orders in an Exact Soluble Model,
Phys. Rev. Lett. 90, 016803 (2003).
[38] A. Kitaev, Anyons in an Exactly Solved Model and Beyond,
Ann. Phys. (Amsterdam) 321, 2 (2006).
[39] J. C. Y. Teo, A. Roy, and X. Chen, Unconventional Fusion
and Braiding of Topological Defects in a Lattice Model,
Phys. Rev. B 90, 115118 (2014).
[40] D. Van Harlingen, private communication.
PHYS. REV. X 5, 041038 (2015)
[41] B. van Heck, F. Hassler, A. R. Akhmerov, and C. W. J.
Beenakker, Coulomb Stability of the 4π-Periodic Josephson
Effect of Majorana Fermions, Phys. Rev. B 84, 180502(R)
(2011).
[42] R. Hutzen, A. Zazunov, B. Braunecker, A. L. Yeyati, and R.
Egger, Majorana Single-Charge Transistor, Phys. Rev.
Lett. 109, 166403 (2012).
[43] B. van Heck, A. R. Akhmerov, F. Hassler, M. Burrello, and
C. W. J. Beenakker, Coulomb-Assisted Braiding of
Majorana Fermions in a Josephson Junction Array, New
J. Phys. 14, 035019 (2012).
[44] J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I. Schuster,
J. Majer, A. Blais, M. H. Devoret, S. M. Girvin, and R. J.
Schoelkopf, Charge-Insensitive Qubit Design Derived from
the Cooper Pair Box, Phys. Rev. A 76, 042319 (2007).
[45] S. Vijay and L. Fu, Physical Implementation of a Majorana
Fermion Surface Code for Fault-Tolerant Quantum
Computation, arXiv:1509.08134.
[46] M. Veldhorst, M. Snelder, M. Hoek, T. Gang, V. K. Guduru,
X. L. Wang, U. Zeitler, W. G. van der Wiel, A. A. Golubov,
H. Hilgenkamp, and A. Brinkman, Josephson Supercurrent
through a Topological Insulator Surface State, Nat. Mater.
11, 417 (2012).
[47] J. R. Williams, A. J. Bestwick, P. Gallagher, S. S. Hong, Y.
Cui, A. S. Bleich, J. G. Analytis, I. R. Fisher, and D.
Goldhaber-Gordon, Unconventional Josephson Effect in
Hybrid Superconductor-Topological Insulator Devices,
Phys. Rev. Lett. 109, 056803 (2012).
[48] F. Yang, F. Qu, J. Shen, Y. Ding, J. Chen, Z. Ji, G. Liu, J.
Fan, C. Yang, L. Fu, and L. Lu, Proximity-Effect-Induced
Superconducting Phase in the Topological Insulator
Bi2 Se3 , Phys. Rev. B 86, 134504 (2012).
[49] J. B. Oostinga, L. Maier, P. Schuffelgen, D. Knott, C. Ames,
C. Brune, G. Tkachov, H. Buhmann, and L. W. Molenkamp,
Josephson Supercurrent through the Topological Surface
States of Strained Bulk HgTe, Phys. Rev. X 3, 021007
(2013).
[50] I. Sochnikov, A. J. Bestwick, J. R. Williams, T. M. Lippman,
I. R. Fisher, D. Goldhaber-Gordon, J. R. Kirtley, and K. A.
Moler, Direct Measurement of Current-Phase Relations in
Superconductor/Topological
Insulator/Superconductor
Junctions, Nano Lett. 13, 3086 (2013).
[51] S. Cho, B. Dellabetta, A. Yang, J. Schneeloch, Z. Xu, T.
Valla, G. Gu, M. J. Gilbert, and N. Mason, Symmetry
Protected Josephson Supercurrents in Three-Dimensional
Topological Insulators, Nat. Commun. 4, 1689 (2013).
[52] C. Kurter, A. D. K. Finck, P. Ghaemi, Y. S. Hor, and D. J.
Van Harlingen, Dynamical Gate-Tunable Supercurrents in
Topological Josephson Junctions, Phys. Rev. B 90, 014501
(2014).
[53] J. D. Sau, R. M. Lutchyn, S. Tewari, and S. Das Sarma,
Generic New Platform for Topological Quantum Computation Using Semiconductor Heterostructures, Phys. Rev.
Lett. 104, 040502 (2010).
[54] J. Alicea, Majorana Fermions in a Tunable Semiconductor
Device, Phys. Rev. B 81, 125318 (2010).
[55] A. P. Higginbotham, S. M. Albrecht, G. Kirsanskas, W.
Chang, F. Kuemmeth, P. Krogstrup, T. S. Jespersen, J.
Nygard, K. Flensberg, and C. M. Marcus, Parity Lifetime
041038-16
MAJORANA FERMION SURFACE CODE FOR UNIVERSAL …
[56]
[57]
[58]
[59]
[60]
[61]
of Bound States in a Proximitized Semiconductor Nanowire,
arXiv:1501.05155.
B. M. Terhal, F. Hassler, and D. P. DiVincenzo, From
Majorana Fermions to Topological Order, Phys. Rev. Lett.
108, 260504 (2012).
Z. Nussinov, G. Ortiz, and E. Cobanera, Arbitrary
Dimensional Majorana Dualities and Architectures for
Topological Matter, Phys. Rev. B 86, 085415 (2012).
C. K. Chiu, D. I. Pikulin, and M. Franz, Strongly Interacting
Majorana Fermions, Phys. Rev. B 91, 165402 (2015).
D. S. Wang, A. G. Fowler, and L. C. L. Hollenberg, Surface
Code Quantum Computing with Error Rates over 1%, Phys.
Rev. A 83, 020302(R) (2011).
D. S. Wang, A. G. Fowler, A. M. Stephens, and L. C. L.
Hollenberg, Threshold Error Rates for the Toric and
Surface Codes, Quantum Inf. Comput. 10, 456 (2010).
F. Hassler, A. R. Akhmerov, and C. W. J. Beenakker,
The Top-Transmon: A Hybrid Superconducting Qubit for
[62]
[63]
[64]
[65]
[66]
041038-17
PHYS. REV. X 5, 041038 (2015)
Parity-Protected Quantum Computation, New J. Phys. 13,
095004 (2011).
L. Sun, L. DiCarlo, M. D. Reed, G. Catelani, Lev S. Bishop,
D. I. Schuster, B. R. Johnson, G. A. Yang, L. Frunzio, L.
Glazman, M. H. Devoret, and R. J. Schoelkopf, Measurements of Quasiparticle Tunneling Dynamics in a Band-GapEngineered Transmon Qubit, Phys. Rev. Lett. 108, 230509
(2012).
A. G. Fowler, Low-Overhead Surface Code Logical Hadamard, Quantum Information Computation 12, 970 (2012).
J. Edmonds, Paths, Trees, and Flowers, Can. J. Math. 17,
449 (1965).
A. G. Fowler, A. C. Whiteside, and L. C. L. Hollenberg,
Towards Practical Classical Processing for the Surface
Code, Phys. Rev. Lett. 108, 180501 (2012).
L. Jiang, C. L. Kane, and J. Preskill, Interface between
Topological and Superconducting Qubits, Phys. Rev. Lett.
106, 130504 (2011).
Download